Previously, we defined eigenvectors and eigenvalues, and used them to obtain a succinct formula for the -th Fibonacci number
, known as Binet’s formula:
where denotes the golden ratio. We needed to take advantage of the eigenvectors and eigenvalues of the matrix
. In particular, we needed to find two eigenvectors
that can encode any
. That is, we wanted a basis for
comprised of the eigenvectors of
.
Let be a field and
be a vector space over
.
Definition 1. Let be a vector space over
. The linear transformation
is similar to
if there exists an isomorphism
such that
. Equivalently,
. We remark that the “is similar to” relation behaves like an equivalence relation.
Example 1. If is finite dimensional with
, then any linear transformation
is similar to some linear transformation
. This means that computations in
can effectively be reduced to computations in
.
Lemma 1. Let and
be similar linear transformations. Then there exists an isomorphism
such that for any
,
where denotes
-fold composition of
.
Roughly speaking, Definition 1 gives us a condition on which and
have essentially the same (i.e. similar) effect on the vectors in their respective vector spaces. The similarity vocabulary empowers us to discuss diagonalisation.
Definition 2. A matrix is diagonal if each
is an eigenvector of
, i.e. for each
, there exists
such that
.
In matrix notation, we have
In particular, for any ,
. A matrix
is diagonalisable if it is similar to some diagonal matrix.
Theorem 1. A matrix is diagonalisable if and only if there exists a basis
of
such that each
is an eigenvector of
.
Proof. We prove this theorem in two directions. Suppose is diagonalisable. Then there exists a diagonal matrix
and an invertible matrix
such that
Writing and employing the recipe-ingredient analogy for matrix multiplication,
Since for each
,
forms the desired basis proving the direction
. Reversing the argument is not hard, and establishes the direction
.
One theoretical consequence is that diagonalisation only proves meaningful when discussing vector spaces that have convenient “standard bases”, such as with the standard ordered basis
. Nevertheless, the discussion on eigenstuff generally still works in broad contexts, especially if the vector space is finite-dimensional.
Corollary 1. Suppose . A linear transformation
is diagonalisable if it is similar to a diagonalisable matrix
. Then
is diagonalisable if and only if there exists a basis
of
such that each
is an eigenvector of
.
Definition 3. A linear transformation is diagonalisable if there exists a basis
of
such that each
is an eigenvector of
.
Therefore, we can reduce our study of general finite-dimensional vector spaces to the concrete setting of . In this case, what sufficiently simple tests do we have to know if a matrix
is diagonalisable?
Lemma 2. Any matrix has at most
distinct eigenvalues. In particular, if
has
distinct eigenvalues, then
is diagonalisable.
Proof. Each eigenvalue yields an eigenvector
, and since distinct eigenvalues yield linearly independent eigenvectors
, the set
is linearly independent, so that
. In particular, if
, the set forms a basis for
, as required.
Example 1. Since has two eigenvalues
, it is diagonalisable, and by Corollary 1, we could extract two key vectors
such that
trivially. This property was key in helping us derive the formula for the Fibonacci numbers.
Now, what if we had less than eigenvalues? Is all hope lost? Not quite. Theorem 1 only requires us to have
eigenvectors. Some of them might have the same eigenvalues—that’s alright. Let
denote the
eigenvalues of
. Each eigenvalue
yields at least one nonzero eigenvector
such that
It is clear that .
Definition 4. We call the function the characteristic polynomial of
. In particular,
is an eigenvalue of
if and only if
.
Lemma 3. If are similar, then
.
Proof. Let be a matrix such that
. Then
By the factor theorem, there exists a polynomial and unique integers
such that
We call the algebraic multiplicity of
. If
, then by the fundamental theorem of algebra, we can take
so that
In any case, the point from Lemma 2 is that captures all of the eigenvectors of
with eigenvalue
. Denote
, called the geometric multiplicity of
. We have a fascinating result.
Theorem 2. For any eigenvalue of
for
,
. Hence, if
and
for all
, then
is diagonalisable.
Proof. Let denote the eigenvectors in
. Extend to a basis
of
with corresponding invertible matrix
. Then
is similar to the matrix
not necessarily diagonal. It follows that
Hence, is a factor of
, which itself contains a factor of
, which yields
. In particular, over
,
is diagonalisable if and only if for each eigenvalue
,
.
Far from just determining eigenvalues, the characteristic polynomial boasts remarkable algebraic properties, encapsulated by the Cayley-Hamilton theorem. In particular, rather surprisingly, we can use characteristic polynomials to compute inverse matrices. We will explore these ideas in the next post.
—Joel Kindiak, 7 Mar 25, 1707H
Leave a comment